• No results found

Trends in analytical separations of magnetic (nano)particles

N/A
N/A
Protected

Academic year: 2022

Share "Trends in analytical separations of magnetic (nano)particles"

Copied!
28
0
0

Laster.... (Se fulltekst nå)

Fulltekst

(1)

Mónica N. Alves1, Manuel Miró2, Michael C. Breadmore1, Mirek Macka1,3,4*

2

3

1 School of Natural Sciences and Australian Centre for Research on Separation Science (ACROSS),

4

University of Tasmania, Hobart 7001, Australia

5

2 FI-TRACE group, Department of Chemistry, University of the Balearic Islands, Carretera de Valldemossa,

6

km. 7.5, E-07122 Palma de Mallorca, Spain

7

3 Department of Chemistry and Biochemistry, Mendel University in Brno, Zemedelska 1, 613 00 Brno,

8

Czech Republic

9

4 Central European Institute of Technology, Brno University of Technology, Purkynova 123, 612 00 Brno,

10

Czech Republic

11

12

*Corresponding author

13

Professor Mirek Macka, Private Bag 75, School of Natural Sciences and Australian Centre for Research on

14

Separation Science, University of Tasmania, Hobart 7001, Australia

15

E-mail: Mirek.Macka@utas.edu.au

16

Phone: +61 362266670; Fax: +61 362262858

17

18

Keywords:

19

Capillary electrophoresis; field flow fractionation; magnetic (nano)particles; magnetophoresis; microfluidic

20

chip; separation

21

22

(2)

Abstract

24

Magnetic particles (MPs) and magnetic nanoparticles (MNPs) are appealing candidates for biomedical and

25

analytical applications due to their unique physical and chemical properties. Given that magnetic fields can

26

be readily used to control the motion and properties of M(N)Ps, their integration in analytical methods

27

opens new avenues for sensing and quantitative analysis. There is a large body of literature related to their

28

synthesis, with a relatively small number of methods reporting the analysis of M(N)Ps using separation

29

methods, which provide information on their purity and monodispersity. This review discusses analytical

30

separation methods of M(N)Ps published between 2013 and June 2018. The analytical separation methods

31

evaluated in this work include (i) field flow fractionation, (ii) capillary electrophoresis, (iii) macroscale

32

magnetophoresis and (iv) microchip magnetophoresis. Among the trends in analytical separations of

33

M(N)Ps an inclination towards miniaturization is moving from conventional benchtop methods to rapid and

34

low-cost methods based on microfluidic devices.

35

(3)

1 Introduction……….…4

37

2 Separations of magnetic (nano)particles……….5

38

2.1 Field flow fractionation………5

39

2.1.1 Magnetic field flow fractionation……….…5

40

2.1.1.1 Magnetic quadrupole field flow fractionation……….6

41

2.1.2 Asymmetrical flow field flow fractionation………8

42

2.1.3 Cyclical electrical field flow fractionation………8

43

2.2 Capillary electrophoresis………9

44

2.3 Macroscale magnetophoresis………...……….…………..……….10

45

2.3.1 High gradient magnetic separation……….………10

46

2.3.2 Low gradient magnetic separation…….……….11

47

2.4 Microchip magnetophoresis……….12

48

3 Conclusion and outlook……….………18

49

4 Acknowledgements……….18

50

5 References……….………19

51

(4)

4 52

1 Introduction

53

Magnetic (nano)particles (M(N)Ps) offer the unique advantage of being manipulated (moved or

54

held in place) using permanent magnets or electromagnets, a significant reason behind their popularity,

55

which grew rapidly in the past decades [1-3]. As an example, one of the most routinely used methods

56

exploiting MPs is magnetic sorting of cell populations from biological suspensions [4]. This method is now

57

standardized for tissue engineering and medical analysis.

58

Magnetic nanoparticles (MNPs) exhibit physical properties that differ remarkably from those of

59

the bulk ferromagnetic material due to finite size effects such as high surface-to-volume ratio, and a special

60

magnetic property at diameters typically lower than 20 nm called superparamagnetism [5]. At such small

61

size, MNPs do not exhibit multiple magnetic domains like ferromagnetic particles, but instead a single

62

domain. Therefore, under an external magnetic field, the magnetic moment of single domain nanoparticles

63

quickly aligns with the applied field, but in its absence, they exhibit no net magnetisation due to the rapid

64

reversal of their magnetic moment. It makes superparamagnetic nanoparticles especially suitable when

65

looking for fast responses to external magnetic fields without agglomeration effects [6]. For this reason,

66

superparamagnetic nanoparticles have been extensively pursued for a vast variety of biomedical

67

applications, including biosensing [7], bioanalysis [8], drug delivery [9], magnetic resonance imaging (MRI)

68

[10], and hyperthermia treatment of tumours [11].

69

The unique behavior and increasing applications of M(N)Ps have stimulated the advancement of

70

new synthesis methods. The core of M(N)Ps can be made of a wide range of magnetic materials such as

71

nickel, cobalt, iron and iron oxides. Iron oxides and their corresponding ferrites are the most commonly

72

used due to their high magnetic moments, biological compatibility, simple synthesis and low cost of

73

production. However, bare iron oxide nanoparticles are only stable in low ionic strength solutions at pH

74

values above (pH 9 - 12) or below (pH 2 - 5) their point zero charge. To prevent aggregation and increase

75

selectivity, the magnetic cores are usually coated with inorganic materials [12], polymers [13-16] and/or

76

functionalised with organic and biological molecules [1, 17]. There is a large quantity of literature exploiting

77

(5)

5

the synthesis and surface engineering of M(N)Ps for many purposes. Compared to this large output, the

78

body of reports on separation approaches used for the analysis of M(N)Ps is considerably smaller, but still

79

very significant. A search in Elsevier’s database Scopus shows that only ca. 12% of the total publications on

80

M(N)Ps address analytical separation techniques. Yet, the separation and analysis of M(N)Ps is critically

81

important to obtain information on their size, shape and chemistry surface, enabling their practical use for

82

many applications.

83

In 2012, Stephens et al. [18] reviewed 58 papers describing separation of M(N)Ps by means of

84

applied magnetic fields and field gradients for improved purification and analysis [18]. The main goal of this

85

review is to pinpoint separation techniques of M(N)Ps for the period between 2013 and June 2018, by

86

employing three distinct types of field-flow fractionation (FFF) (magnetic FFF, asymmetrical FFF and cyclical

87

electrical FFF), capillary electrophoresis (CE), macroscale magnetophoresis (high/low gradient magnetic

88

separation), and microchip magnetophoresis. Analytical separation and sample preparation approaches

89

using M(N)Ps are deemed outside the scope of our review, for which there is a large body of literature,

90

including recent comprehensive reviews [1-3, 19-23]. All the studies discussed and critically analyzed are

91

summarized in table 1 in terms of particle composition and size, magnetic field applied, separation principle,

92

separation time and complexity of the infrastructures used.

93 94

2 Separations of magnetic (nano)particles

95

2.1 Field flow fractionation

96

FFF is a separation technique that uses an external field applied perpendicular to the direction of flow

97

causing differential migration of M(N)Ps. Typically, the flow profile in a FFF channel is laminar, so particles

98

which interact more strongly with the field are found closer to the channel walls and will move more slowly

99

due to slower flow streams. Analytes can be separated by different mechanisms of FFF according to the

100

type of the field applied. Typical fields include centrifugal and gravitational forces, cross flow of solvent,

101

and thermal, electrical and magnetic gradients [24].

102

103

(6)

6

2.1.1 Magnetic field flow fractionation

104

Magnetic field flow fractionation (MFFF) has been shown to be an effective method for the

105

separation of polydisperse suspensions of M(N)Ps when an external magnetic field is applied along a flowing

106

channel. Rogers et al. [25] used MATLAB to simulate the separation of fluidMAG-D (starch-coated magnetite

107

(Fe3O4)) M(N)Ps of sizes between 50 and 400 nm by simulating particle trajectories and magnetic forces.

108

The results obtained from the simulation showed that M(N)Ps within the size range of interest could be

109

separated and collected in a size dependent manner in fraction 1 (smaller sized) and fraction 2 (larger sized).

110

To validate this model, the simulated conditions were replicated experimentally. However, the theoretical

111

and experimental results did not agree. This inconsistency could be due to the fact that particle-particle

112

interactions are not taken into consideration into the model. Due to the failure of the initial experiments,

113

the same authors changed the approach to a simple magnetic coil setup composed of a tubing wrapped

114

around a Grade N42 diametrically magnetized neodymium cylinder [25]. An inlet for both the M(N)Ps

115

suspension and the mobile phase was inserted at the top of the magnetic coil and a single outlet at the

116

bottom was used for the collection of magnetic particle fractions. The tubing was filled with M(N)Ps

117

suspension. As soon as a steady state level of accumulation of M(N)Ps across the inner wall of the tubing

118

was achieved, an initial flow rate of 0.25 mL/min was applied to wash out the M(N)Ps remaining suspended.

119

Then, the flow rate was increased up to 50 mL/min. DLS measurements and TEM showed that the particles

120

collected at lower flow rates were smaller than particles collected at higher flow rates. This approach was

121

low cost and allowed the separation of polydisperse M(N)Ps by simply controlling the flow rate. However,

122

broad size distributions were obtained. Thus, further optimisation of the system is crucial to allow for

123

specific applications such as in biomedicine.

124 125

2.1.1.1 Magnetic quadrupole field flow fractionation

126

The first prototype of a magnetic quadrupole field flow fractionation (MQFFF) was developed and

127

evaluated by Zborowski et al. [26] for continuous separation of human peripherical lymphocytes labeled

128

with magnetic colloids. It consisted of a quadrupole electromagnet assembly of four steel pole tips with two

129

(7)

7

of them opposed the magnetic north poles and the other two opposed the magnetic south poles. The

130

electromagnet assembly was radially symmetric. The steel was magnetized by an electric current in the coils

131

wrapped around the poles. This configuration creates a magnetic field whose magnitude increases linearly

132

with the radial distance from the axis. This methodology is well described by Carpino et al. [27]. The results

133

showed that the separation process was close to the predicted behavior of an ideal quadrupole magnetic

134

field [26]. Later on, Orita et al. [28] developed a simple on-off field MQFFF to separate and quantify two

135

distinct sub micrometer commercial M(N)Ps (90 and 200 nm) at specific magnetic field and flow conditions

136

[28]. This on-off field MQFFF system was inspired by the system previously developed by Zborowski et al.

137

[26] . It consisted of a separation channel volume of 0.94 mL fitted into a stainless-steel cylinder that was

138

implemented in a flow injection setup with downstream optical detection (Figure 1). The fractograms

139

exhibited improved retention (98.6% vs. 53.3%) for the larger M(N)Ps (200 nm vs.90 nm) at higher flow

140

rates (0.05 mL/min vs. 0.01 mL/min) [28]. Thus, for given field and flow conditions, the on-off field MQFFF

141

system can be used for the quantification of retained and unretained fractions. This is useful for the

142

separation of unwanted weakly magnetic particulate contaminants from M(N)P suspensions. Compared to

143

the magnetic coil setup previously described, the on-off field MQFFF system requires less handling.

144 145

146

Figure 1 – Schematic of the MQFFF system. The mobile phase was driven by a pump. The sample is

147

introduced through a separate port. The flow of the mobile phase pushes the sample from the injector into

148

(8)

8

the separation channel fitted into a quadrupole electromagnet connected to a UV-visible detector.

149

Reprinted from [28] with permission.

150 151

Moore et al. [29] used MQFFF for red blood cell (RBC, with mean diameter of 8 µm) separation as

152

an alternative to centrifugal separation. A quadrupole field was designed having a maximum field of 1.68 T

153

at the magnet pole tips, zero field at the aperture axis, and a nearly constant radial field gradient of 1.75 T

154

mm-1 inside a cylindrical aperture. A light-scattering detector downstream of the magnet measured light

155

attenuation caused by the cells eluting from the magnet as a function of time. The cell samples were

156

composed of high spin RBC (obtained by chemical conversion of hemoglobin to methemoglobin - met RBC

157

- or exposure to anoxic conditions - deoxy RBC), low spin RBC (obtained by exposure to ambient air – oxy

158

RBC), and mixtures of deoxy RBC and white blood cells (WBC). Cell tracking velocimetry was used to

159

measure the magnetophoretic mobility of the RBC and the results showed that the mobility depended on

160

the presence of high-spin hemoglobin. Only high spin RBC were attracted by the magnet, while low-spin

161

RBC demonstrated magnetic susceptibility comparable to WBC. It was also found that RBC did not elute

162

within 15 min from the channel at flow rate ≤ 0.05 mL/min but as would be expected rapidly eluted at a

163

higher flow rate of 2.0 mL/min. These results agreed with earlier studies on the magnetic properties of

164

hemoglobin using other techniques [30, 31]. The fractionation experiments of RBC and a RBC/WBC mixture

165

showed that a 5 x 107/mL cell suspension pumped at 0.1 mL/min through a magnetic field of 1.5T and

166

gradient of 1,000 T m-1 is depleted to less than 5% of the initial RBC number concentration. The 5% residual

167

contamination was comparable to that typically seen in WBC obtained by blood centrifugation. One

168

advantage over blood centrifugation, is that MQFFF RBC separation can be scaled to microliter devices for

169

RBC debulking, which can be portable and operated immediately after donation with minimal human labor.

170 171

2.1.2 Asymmetrical flow field-flow fractionation

172

Asymmetrical flow field flow fractionation (AF4) is a separation technique based on the theory of

173

FFF. The cross flow is induced by flowing liquid constantly exiting through a semi-permeable wall on the

174

(9)

9

bottom of the channel. The lower size M(N)Ps, which can be fractioned, are restricted by the molecular

175

weight cut off membrane. The suitability of AF4 has been shown for the fractionation of magneto polyplexes

176

(with mean diameter of 54 nm) [32] and carboxydextran-coated maghemite dispersions (with diameter of

177

6 – 60 nm) [33] by connecting the AF4 instrument to UV [33] and multi-angle laser light scattering (MALLS)

178

[32, 33] detectors. AF4 can be a simple, fast and reliable tool for quality control in commercial production

179

of M(N)Ps, while providing complementary information related to nonmagnetic sample components.

180 181

2.1.3 Cyclical electrical field flow fractionation

182

Cyclical electrical field flow fractionation (CyEFFF) consists of an oscillating square voltage applied

183

between a top and bottom electrode inside the channel. As result, M(N)Ps move back and forth between

184

the electrodes in agreement with their sizes and electrophoretic mobilities. M(N)Ps with high

185

electrophoretic mobilities move further into the center of the channel and they spend more time at the

186

faster fluid regions thus eluting earlier than the lower mobility M(N)Ps. One of the limitations of this

187

technique is the band broadening of the resulting UV fractograms and low resolution. To address and solve

188

the diffusion issue, Tasci et al. [34] reported the separation of MNPs by CyEFFF applying square wave

189

voltages with higher duty cycles instead of DC offset voltages (Figure 2). Thus, particle diffusion was

190

suppressed, which allowed separations of MNPs with mean diameter of 50 and 100 nm. This study

191

demonstrated the capability of CyEFFF for size and electrophoretic analysis of lipid and polystyrene

192

sulfonate-coated MNPs [34].

193

194

Figure 2 – Schematic of the CyEFFF system. The dashed line shows the particle trajectory that results from

195

the cyclical electrical field. Reprinted from [34] with permission.

196

(10)

10

2.2 Capillary electrophoresis

197

CE is a powerful separation method which has the advantages of minimal requirement of sample and

198

buffer volumes, and lack of generation of organic waste. Further, the use of narrow capillaries in CE with

199

high electrical resistance, allows the application of high electrical fields with minimal heat generation. The

200

use of high electrical fields together with the conventional plug-type flow from electrically driven systems

201

results in short analysis times and high efficiency and resolution for almost any type of ionic analytes. Yet,

202

a remarkable limitation of bare iron oxide M(N)P separations by CE that has been previously reported [35-

203

38] is their high tendency to spontaneously agglomerate to minimize surface energies, which is observed in

204

electropherograms as spurious spikes that in turn prevent the accurate determination of the

205

electrophoretic mobilities of M(N)Ps. This limitation has been recently overcome by Alves et al. [39], who

206

achieved symmetrical and smooth peaks of bare iron oxide nanoparticles. This was accomplished through

207

electrostatic stabilisation using complexing electrolyte anions such as citrate and phosphate, and the

208

additive tetramethylammonium hydroxide (TMAOH) within the background electrolyte (BGE), an ionic

209

solution of desired concentration, co-ion and counter-ion mobilities, and usually also providing pH-buffering

210

capacity. TMAOH is a peptizing agent (an electrolyte that converts aggregated particles into a colloidal sol)

211

[40] used for more than two decades in the synthesis of well-dispersed iron oxide M(N)P solutions, however

212

never utilised for effective CE separations of M(N)Ps. The same study also showed the successful separation

213

of bare (with diameter between 7 and 13 nm) and carboxylated (10 nm) iron oxide nanoparticles in 12 min

214

using Tris-nitrate containing 20 mM TMAOH as BGE. The electrophoretic mobilities for bare and

215

carboxylated iron oxide nanoparticles were 3.3E-08 m2 V-1 s-1 (0.9 %RSD) and 4.1E-08 m2 V-1 s-1 (0.4 %RSD),

216

respectively. These findings demonstrate that simple and rapid CE experiments are excellent tools to

217

characterise and monitor properties and interactions of iron oxide nanoparticles with other molecules for

218

potential surface modification purposes [39]. Baron et al. [41] studied the online stacking of carboxylated

219

core-shell magnetite nanoparticles in CE. By monitoring the ionic strength of the BGE and the sample zone,

220

it was observed that stacking occurred optimally when MNPs were dispersed in 10 mM borate/NaOH (pH

221

9.5) and injected to the BGE composed of 100 mM borate/NaOH (pH 9.5). The decrease of the electric

222

(11)

11

double layer thickness with increasing ionic strength could induce MNP aggregation and led to the

223

restructuring of the MNPs zone due to the decrease of distance between nanoparticles [41]. The Derjaguin-

224

Landau-Verwey-Overbeek (DLVO) theory, which describes van-der-Waals and electrostatic interactions

225

between charged surfaces within a liquid medium [42], was suggested as a cause of peak sharpening in CE.

226 227

2.3 Macroscale magnetophoresis

228

Magnetophoresis refers to the motion of magnetic particles or magnetizable material through a

229

fluid under the influence of a magnetic field [43]. For almost all the M(N)Ps applications, manipulation,

230

recovery, and collection rates using external magnets should be done quickly. Rapid magnetophoretic

231

separation can be attained under both high gradient (HGMS, ∇B⃗ > 1000 T m-1) and low gradient magnetic

232

separation (LGMS, ∇B⃗ < 100 T m-1).

233 234

2.3.1 High gradient magnetic separation

235

HGMS is commonly employed in conventional industry practice to separate magnetic materials from

236

non-magnetic aqueous solutions, such as for wastewater treatment of bacteria and solids. Typically, HGMS

237

is used to separate microscale or bigger particles, or microscale aggregates of nanoparticles, or

238

nanoparticles encapsulated in larger polymer beads. However, the application of HGMS to suspensions of

239

individually dispersed M(N)Ps has been poorly explored so far. HGMS systems generally consist of a column

240

packed with magnetically susceptible wires placed inside an electromagnet. Through application of a

241

magnetic field across the column, the wires dehomogenise the magnetic field in the column producing high

242

field gradients around the wires to enable the capture of M(N)Ps onto their surfaces. The attraction of

243

M(N)Ps depends on the magnetic field gradients generated, particle size and magnetic properties [44]. A

244

study conducted by Mirshahghassemi et al. [45] describes the application of HGMS for oil remediation using

245

polyvinylpyrrolidone (PVP)-coated MPs (mean size diameter of 127 nm) in a continuous and large volume

246

flow system. This technique was analyzed as a function of magnetic field strength, mixing time and stainless-

247

(12)

12

steel wool content. Fluorescence and inductively coupled plasma-optical emission spectrometer (ICP-OES)

248

data indicated that ca. 85% of oil and 95% of MNP were eliminated. The continuous use of this HGMS system

249

over 7 hours allowed the treatment of 17 liter oil water mixture with no reduction of the oil and MPs

250

removal capacity [45]. Although this study introduces a new application of HGMS for oil remediation, it has

251

the disadvantages of being tedious and time-consuming.

252 253

2.3.2 Low gradient magnetic separation

254

In contrast to conventional practice where HGMS is normally employed, LGMS is still poorly explored

255

and understood. Toh et al. [46, 47] showed the reliability of LGMS for magnetophoretic separation of

256

microalgal biomass that interacted electrostatically with cationic polymer functionalized MNPs [46, 47].

257

Poly (diallyl dimethylammonium chloride) (PDDA) and chitosan (Chi) (with mean diameter of 50 nm) worked

258

as binding agents to promote rapid separation of the negatively charged Chlorella sp. through LGMS at field

259

gradient lower than 80 T m-1. The obtained results indicated cell separation efficiency of about 98% for

260

PDDA and 99% for Chi. Though, from a practical point of view, PDDA was preferable as polymer binder since

261

the attachment mechanism involved was pH independent [47]. Because almost all the magnetophoretic

262

studies have been dedicated to the behavior of spherical MNPs, and poor attention has been paid to rod-

263

like MNPs, Lim et al. [48] compared the magnetophoretic behavior of spherical and rod-like iron oxide

264

nanoparticles under LGMS. Both effects of particle concentration and magnetic field gradient on the

265

separation kinetics were evaluated. It was shown that at low particle concentration, the magnetophoresis

266

of MNPs at low magnetic gradient is significantly enhanced by particle anisotropy (non-spherical shape),

267

with rod-like MNPs taking significantly less time than spherical MNPs to be separated [48]. New approaches

268

for the separation of M(N)Ps according to their shape, size, and coatings, along with their unique magnetic

269

properties will open new opportunities for M(N)Ps.

270 271

2.4 Microchip magnetophoresis

272

(13)

13

The field of microfluidics is continuously evolving as miniaturized platforms provide quick analysis

273

with high resolution at low cost, foster portability and make use of exceptionally minute amounts of

274

reagents. Zhang et al. [49] developed a microchip based on magnetophoresis with differential interference

275

contrast (DIC) detection (Figure 3A and 3B). The real-time moving trajectories and velocities of the MPs at

276

different magnetic field strengths (depending on the distance to the permanent magnet) were measured

277

based on consecutive DIC images. The results indicated that shorter distances to the magnet caused higher

278

magnetophoretic velocities of MPs, and enhanced magnetophoretic velocity differences between dissimilar

279

particle sizes (Figure 3D) [49]. This study allowed the successful separation and detection of a polydisperse

280

mixture of MPs (500 nm and 160 nm) at a single-particle level in only about 15 seconds (Figure 3C).

281

282

Figure 3 - Schematic diagram (A) and photography (B) of the experimental setup of the microchip

283

magnetophoresis. (C) Representative magnetopherograms of the MPs by microchip magnetophoresis with

284

the DIC detection system. (D) Magnetophoretic velocities of the MPs at different permanent magnet

285

distances obtained using a DIC microscope. Reprinted from [49] with permission.

286 287

The microfluidic separation of iron oxide beads (5 µm diameter) with soft magnetic

288

microstructures was demonstrated by Zhou et al. [50]. The fabrication process of the microfluidic device is

289

represented in Figure 4. This microfluidic device consisted of two channels - fluidic and structural – made

290

in polydimethylsiloxane (PDMS). The fluidic channel contained two inlets and two outlets. A mixture of iron

291

(14)

14

powder and PDMS was injected into the structural channel located between two external permanent

292

magnets. Three microstructure shapes were studied: (i) half circle, (ii) 60o isosceles triangle and (iii) 120 o

293

isosceles triangle. The soft magnetic microstructures provided localized and strong magnetic forces on the

294

MPs that deflected them perpendicularly to the pressure-driven flow. Thus, the separation depended on

295

the magnetic forces. In turn, magnetic forces are affected by the shape of the iron-PDMS microstructures

296

and the mass ratio of the iron-PDMS composite. Also, the flow rate in the fluid channel affects the time that

297

MPs are subjected to the magnetic field, and consequently their vertical deflection. Numerical simulations

298

were developed to predict the particle trajectories showing good agreement with experimental data. Finally,

299

systematic experiments and simulations were conducted to study the effect of several relevant factors on

300

the separation of MPs: microstructure shape, mass ratio of the iron–PDMS, microfluidic channel width and

301

average flow velocity. The results demonstrated that (i) half circular iron–PDMS microstructure caused

302

greater deflections, (ii) larger mass ratio of the iron–PDMS composite provided higher magnetic forces, and

303

(iii) wider channels separate MPs less efficiently than narrow microfluidic channels when operating at the

304

same flow rate [50]. Based on the obtained results, enhanced separations of MPs can be achieved in a

305

compact, simple and low-cost microfluidic device.

306 307

308

(15)

15

Figure 4 - Fabrication process of the microfluidic device for efficient separation of magnetic particles based

309

on deflection in flowing streams. Reprinted from[50] with permission.

310 311

Kumar and Rezai [51] introduced a novel hybrid technique called multiplex inertio-magnetic

312

fractionation (MIMF) to simultaneously fractionate up to four magnetic and non-magnetic particles in water

313

at a throughput of 106 – 109 particles per hour. The MIMF device was based on interactions between flow-

314

induced inertial forces and magnetic forces in an expansion microchannel containing an external permanent

315

magnet (Figure 5). The particle fractionation performance was first optimized in terms of flow rate and

316

aspect ratio of the channel and particle size on duplex MIMF to understand the behavior of the particles.

317

The obtained knowledge was then applied to demonstrate fourplex MIMF with three magnetic

318

monodisperse particles (5, 11 and 35 µm) and nonmagnetic particles (15 µm). The non-magnetic particles

319

inertially focus at the center of the channel, while magnetic particles get fractionated based on interaction

320

between inertial and magnetic forces and positioned in a size related manner in the device with smaller

321

particles located closer to the external magnet. The exit position for each particle type and size was

322

measured with respect to the expansion region baseline since the focus of this study was only to investigate

323

the concept of MIMF and not M(N)Ps sorting. In the future, outlets can be implemented based on the exit

324

positions calculated. This MIMF device addresses several disadvantages of currently available magnetic

325

fractionation devices such as low throughput, requirement of sheath flow and inability to fractionate

326

multiple targets simultaneously [51].

327

(16)

16 328

Figure 5 – MIMF scheme of the particle separation. The device (scale bar 25 mm) consisted of an inertio-

329

magnetic zone (IMZ) with a side permanent magnet and an expansion zone (EZ). The schematic

330

representation of MIMF device shows three red-colored magnetic particles (MP) of varied sizes and a black-

331

colored non-magnetic particle (NMP). Reprinted from [51] with permission.

332 333

Dutz et al. [52] studied the consequences of applying an external magnetic force to a suspension

334

of MPs with diameters of 2, 6 and 12 µm circulating in a spiral microfluidic channel (Figure 6). The fluid was

335

injected via an input port located near the center of the spiral and exits through a symmetric flow splitter

336

and two outlet ports. The exit ports were referred to as inner outlet and outer outlet, which collected the

337

inner and outer halves of the fluid stream, respectively. For that purpose, an array of permanent magnets

338

was arranged and accurately centered beneath the spiral to produce a magnetic field with octupolar

339

symmetry. At low flow rates (5 µL/min) it was observed that 6 µm MPs clustered along a streamline near

340

the outer wall of the spiral. At intermediate flow rates (30 µL/min), 6 µm MPs were homogeneously

341

distributed across the width of the channel. At high flow rates (60 µL/min), 6 µm MPs were focused in

342

clusters within the inner half of the spiral. The phenomenon observed at high flow rates was caused by

343

hydrodynamic drag forces that induced secondary (Dean) flow in the spiral microfluidic channel. A model

344

(17)

17

incorporating key forces involved in the spiral microchip was described and used to extract quantitative in

345

situ information about the magnitude of local Dean drag forces from experimental data. The experimental

346

results also showed that at low flow rates (5 µL/min) all the 12 μm MPs and one third of the 2 μm MPs were

347

drawn toward the outer wall of the spiral and extracted from the outer outlet, whereas the remaining two

348

thirds of the 2 μm MPs were extracted from the inner outlet. Gradually more 6 and 12 μm MPs were

349

extracted from the inner outlet at higher flow rates. For example, all the 12 μm MPs were found to exit the

350

inner outlet at 40 µL/min. The behavior of the smallest MPs (2 μm) was opposite of the larger particles,

351

with less and less being extracted from the inner outlet as the flow rate increases. [52].

352

353

Figure 6 – Geometry of the microfluidic spiral. The fluid was injected via the input port (in) and exits through

354

the exit ports referred as inner outlet (io) and outer outlet (oo). Reprinted from [52] with permission.

355 356

The effective application of MNPs is highly dependent of appropriate cleaning after synthesis

357

and/or before their use to remove solvents, excess of surfactants, byproducts and undesired impurities.

358

Because manual cleaning is time consuming and inefficient, Cardoso et al. [53] designed, fabricated and

359

tested a microfluidic system for the continuous cleaning and separation of MNPs (average diameter of 10

360

nm) synthesized by coprecipitation using NH4OH as catalyst. First, a theoretical study was performed to

361

optimize the geometrical configuration of the microfluidic device and the experimental conditions. The

362

optimized microfluidic system was composed of two inlets and two outlets (Figure 7). The cleaning solution

363

(water, fluid A) was introduced through the inlet A and the synthesis solution with MNP (fluid B) through

364

the inlet B. The waste fluid (fluid C) exited through the outlet C, whereas the cleaned MNPs solution (fluid

365

(18)

18

D) exited though the outlet D. A permanent magnet was located near the diffusion channel to deflect the

366

MNPs by magnetic forces from the synthesis solution to the cleaning solution (figure 7). Gas

367

chromatography was performed to indirectly calculate the cleaning efficiency by measuring the decrease

368

of the peak area of NH4OH. The results demonstrated a cleaning efficiency of about 99.7% by controlling

369

the fluid flows in the microfluidic system, whereas manual cleaning achieved a value of about 94.3% after

370

cleaning six consecutive times. Both processes are time-consuming, however the microfluidic system offers

371

negligible loss of MNPs and the process is performed autonomously [53].

372 373 374

375

Figure 7 – Schematic of the optimized microfluidic system. Fluid A: cleaning solution; Fluid B: synthesis

376

solution with MNPs; Fluid C: waste; Fluid D: cleaned MNPs. Channel widths (a) 600 µm; (b) 400 µm; (c) 600

377

µm; (d) 400 µm; diffusion channel length (e) 10 mm. Reprinted from [53] with permission.

378 379

The separation of magnetic particles with spherical (mean diameter of 7 µm) and elliptical shapes

380

was demonstrated by Zhou et al. [54] in a simple and effective manner. The microfluidic chip consisted of

381

two inlets and one outlet. The inlet 1 was injected with aqueous-glycerol solution that worked as buffer

382

flow, while the inlet 2 was injected with sample particles suspended in aqueous-glycerol solution. The

383

microfluidic device was placed in the center of a uniform magnetic field and mounted on an inverted

384

microscope to record the trajectories of the magnetic particles. A pressure-driven flow was combined with

385

the magnetic field applied perpendicularly to the flow direction. The results showed that the asymmetrical

386

rotation of the ellipsoidal MPs, together with the particle-wall hydrodynamic interactions, resulted in a net

387

(19)

19

lift force towards the channel center. Differently, spherical MPs remained closer to the channel wall. This

388

uniform magnetic field technique can be applied to multiple microfluidic channels facilitating high

389

throughput parallelization for biological and biomedical applications that require separation of shaped MPs

390

[54].

391 392

Conclusion and outlook

393

The improvement of the existing approaches of synthetizing uniform and more monodisperse

394

M(N)Ps has been notorious in recent years. However, even the most efficient and highly optimized

395

protocols yield samples relatively polydisperse. This is an obstacle for some emerging applications of

396

M(N)Ps with some specific functions, particularly in biomedical areas, as their properties are dependent.

397

Analytical tools for M(N)Ps separation are fundamental to understand the behavior of M(N)Ps according to

398

their size, shape and surface chemistry. This knowledge can give information on the monodispersity and

399

purity of M(N)Ps for their ultimate practical use. Over the last 5 years, novel strategies for M(N)P

400

separations have been reported based on FFF, macroscale magnetophoresis and CE, but the trend is toward

401

integrating external magnetic fields onto single microfluidic structures. From about 33 journal articles that

402

have been published in the last 5 years, 54% of them are microfluidic based. These can be easily fabricated

403

and are inexpensive, not requiring any extra power source. Moreover, separations in microchip

404

magnetophoresis can be easily achieved with high efficiency and throughput in the order of seconds. Real-

405

time moving trajectories and velocities of M(N)Ps can also be monitored by means of microscope imaging

406

that is seen as the new trend in microfluidic separation and identification of M(N)Ps.

407

408

409

(20)

20

Acknowledgements

410

Mirek Macka gratefully acknowledges the Australian Research Council Future Fellowship (FT120100559).

411

Manuel Miró acknowledges financial support from the Spanish State Research Agency (AEI) through project

412

CTM2017-84763-C3-3-R (MINECO/AEI/FEDER, EU). Michael Breadmore acknowledges an Australian

413

Research Council Future Fellowship Award (FT130100101).

414

415

(21)

21

References

416

[1] Á. Ríos, M. Zougagh, Recent advances in magnetic nanomaterials for improving analytical

417

processes, TrAC, Trends Anal. Chem. 84 (2016) 72-83.

418

[2] T. Jamshaid, E.T.T. Neto, M.M. Eissa, N. Zine, M.H. Kunita, A.E. El-Salhi, A. Elaissari, Magnetic

419

particles: From preparation to lab-on-a-chip, biosensors, microsystems and microfluidics applications, TrAC,

420

Trends Anal. Chem. 79 (2016) 344-362.

421

[3] I. Vasconcelos, C. Fernandes, Magnetic solid phase extraction for determination of drugs in

422

biological matrices, TrAC, Trends Anal. Chem. 89 (2017) 41-52.

423

[4] S. Miltenyi, W. Müller, W. Weichel, A. Radbruch, High gradient magnetic cell separation with MACS,

424

Cytometry A 11 (1990) 231-238.

425

[5] S.A. Wahajuddin, Superparamagnetic iron oxide nanoparticles: magnetic nanoplatforms as drug

426

carriers, Int. J. Nanomedicine 7 (2012) 3445.

427

[6] S.R. Dave, X. Gao, Monodisperse magnetic nanoparticles for biodetection, imaging, and drug

428

delivery: a versatile and evolving technology, Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 1 (2009)

429

583-609.

430

[7] J.M. Perez, T. O'Loughin, F.J. Simeone, R. Weissleder, L. Josephson, DNA-based magnetic

431

nanoparticle assembly acts as a magnetic relaxation nanoswitch allowing screening of DNA-cleaving agents,

432

J. Am. Chem. Soc. 124 (2002) 2856-2857.

433

[8] J.E. Smith, L. Wang, W. Tan, Bioconjugated silica-coated nanoparticles for bioseparation and

434

bioanalysis, TrAC, Trends Anal. Chem. 25 (2006) 848-855.

435

[9] N. Kohler, C. Sun, J. Wang, M. Zhang, Methotrexate-modified superparamagnetic nanoparticles

436

and their intracellular uptake into human cancer cells, Langmuir 21 (2005) 8858-8864.

437

[10] B. Chertok, B.A. Moffat, A.E. David, F. Yu, C. Bergemann, B.D. Ross, V.C. Yang, Iron oxide

438

nanoparticles as a drug delivery vehicle for MRI monitored magnetic targeting of brain tumors, Biomaterials

439

29 (2008) 487-496.

440

(22)

22

[11] K. Maier-Hauff, F. Ulrich, D. Nestler, H. Niehoff, P. Wust, B. Thiesen, H. Orawa, V. Budach, A. Jordan,

441

Efficacy and safety of intratumoral thermotherapy using magnetic iron-oxide nanoparticles combined with

442

external beam radiotherapy on patients with recurrent glioblastoma multiforme, J. Neuro-Oncol. 103 (2011)

443

317-324.

444

[12] V. Salgueiriño‐Maceira, M.A. Correa‐Duarte, M. Spasova, L.M. Liz‐Marzán, M. Farle, Composite

445

silica spheres with magnetic and luminescent functionalities, Adv. Funct. Mater. 16 (2006) 509-514.

446

[13] G.S. Demirer, A.C. Okur, S. Kizilel, Synthesis and design of biologically inspired biocompatible iron

447

oxide nanoparticles for biomedical applications, J. Mater. Chem. B 3 (2015) 7831-7849.

448

[14] J. Chomoucka, J. Drbohlavova, D. Huska, V. Adam, R. Kizek, J. Hubalek, Magnetic nanoparticles and

449

targeted drug delivering, Pharmacol. Res. 62 (2010) 144-149.

450

[15] A.K. Gupta, M. Gupta, Synthesis and surface engineering of iron oxide nanoparticles for biomedical

451

applications, Biomaterials 26 (2005) 3995-4021.

452

[16] D. Singh, J.M. McMillan, X.-M. Liu, H.M. Vishwasrao, A.V. Kabanov, M. Sokolsky-Papkov, H.E.

453

Gendelman, Formulation design facilitates magnetic nanoparticle delivery to diseased cells and tissues,

454

Nanomedicine 9 (2014) 469-485.

455

[17] W. Wu, Q. He, C. Jiang, Magnetic iron oxide nanoparticles: synthesis and surface functionalization

456

strategies, Nanoscale Res. Lett. 3 (2008) 397.

457

[18] J.R. Stephens, J.S. Beveridge, M.E. Williams, Analytical methods for separating and isolating

458

magnetic nanoparticles, Phys. Chem. Chem. Phys. 14 (2012) 3280-3289.

459

[19] V. Tolmacheva, V. Apyari, E. Kochuk, S. Dmitrienko, Magnetic adsorbents based on iron oxide

460

nanoparticles for the extraction and preconcentration of organic compounds, J. Anal. Chem. 71 (2016) 321-

461

338.

462

[20] Y. Wang, Q. Chen, C. Gan, B. Yan, Y. Han, J. Lin, A review on magnetophoretic immunoseparation,

463

J. Nanosci. Nanotechnol. 16 (2016) 2152-2163.

464

[21] A.A. Hernandez-Hernandez, G.A. Alvarez-Romero, E. Contreras-Lopez, K. Aguilar-Arteaga, A.

465

Castaneda-Ovando, Food Analysis by Microextraction Methods Based on the Use of Magnetic Nanoparticles

466

as Supports: Recent Advances, Food Anal. Methods 10 (2017) 2974-2993.

467

(23)

23

[22] S. Piovesana, A. L. Capriotti, Magnetic materials for the selective analysis of peptide and protein

468

biomarkers, Curr. Med. Chem. 24 (2017) 438-453.

469

[23] T. Tangchaikeeree, D. Polpanich, A. Elaissari, K. Jangpatarapongsa, Magnetic particles for in vitro

470

molecular diagnosis: From sample preparation to integration into microsystems, Colloids Surf. B 158 (2017)

471

1-8.

472

[24] J.C. Giddings, F. Yang, M.N. Myers, Flow-field-flow fractionation: a versatile new separation

473

method, Science 193 (1976) 1244-1245.

474

[25] H.B. Rogers, T. Anani, Y.S. Choi, R.J. Beyers, A.E. David, Exploiting size-dependent drag and

475

magnetic forces for size-specific separation of magnetic nanoparticles, Int. J. Mol. Sci. 16 (2015) 20001-

476

20019.

477

[26] M. Zborowski, L. Sun, L.R. Moore, P.S. Williams, J.J. Chalmers, Continuous cell separation using

478

novel magnetic quadrupole flow sorter, J. Magn. Magn. Mater. 194 (1999) 224-230.

479

[27] F. Carpino, L.R. Moore, J.J. Chalmers, M. Zborowski, P.S. Williams, Quadrupole magnetic field-flow

480

fractionation for the analysis of magnetic nanoparticles, IOP Publishing. 2005, p. 174.

481

[28] T. Orita, L.R. Moore, P. Joshi, M. Tomita, T. Horiuchi, M. Zborowski, A quantitative determination

482

of magnetic nanoparticle separation using on-off field operation of quadrupole magnetic field-flow

483

fractionation (QMgFFF), Anal. Sci. 29 (2013) 761-764.

484

[29] L.R. Moore, P.S. Williams, F. Nehl, K. Abe, J.J. Chalmers, M. Zborowski, Feasibility study of red blood

485

cell debulking by magnetic field-flow fractionation with step-programmed flow, Anal. Bioanal. Chem. 406

486

(2014) 1661-1670.

487

[30] L. Pauling, C.D. Coryell, The magnetic properties and structure of hemoglobin, oxyhemoglobin and

488

carbonmonoxyhemoglobin, Proc. Natl Acad. Sci. U.S.A 22 (1936) 210-216.

489

[31] C.D. Coryell, F. Stitt, L. Pauling, The magnetic properties and structure of ferrihemoglobin

490

(methemoglobin) and some of its compounds, J. Am. Chem. Soc. 59 (1937) 633-642.

491

[32] E. Wetterskog, A. Castro, L. Zeng, S. Petronis, D. Heinke, E. Olsson, L. Nilsson, N. Gehrke, P.

492

Svedlindh, Size and property bimodality in magnetic nanoparticle dispersions: single domain particles vs.

493

strongly coupled nanoclusters, Nanoscale 9 (2017) 4227-4235.

494

(24)

24

[33] E. Haladjova, S. Rangelov, M. Geisler, S. Boye, A. Lederer, G. Mountrichas, S. Pispas, Asymmetric

495

Flow Field‐Flow Fractionation Investigation of Magnetopolyplexes, Macromol. Chem. Phys. 216 (2015)

496

1862-1867.

497

[34] T. Tasci, E. Manangon, D. Fernandez, W. Johnson, B. Gale, Separation of magnetic nanoparticles by

498

cyclical electrical field flow fractionation, IEEE Trans. Magn. 49 (2013) 331-335.

499

[35] G.R. Ducatte, N.E. Ballou, C. Quang, S.L. Petersen, Separation and characterization of oxide

500

particles by capillary electrophoresis, J. Microcolumn Sep. 8 (1996) 403-412.

501

[36] C. Quang, S. Petersen, G. Ducatte, N. Ballou, Characterization and separation of inorganic fine

502

particles by capillary electrophoresis with an indifferent electrolyte system, J. Chromatogr. A 732 (1996)

503

377-384.

504

[37] S.L. Petersen, N.E. Ballou, Separation of micrometer-size oxide particles by capillary zone

505

electrophoresis, J. Chromatogr. A 834 (1999) 445-452.

506

[38] N.G. Vanifatova, B.Y. Spivakov, J. Mattusch, U. Franck, R. Wennrich, Investigation of iron oxide

507

nanoparticles by capillary zone electrophoresis, Talanta 66 (2005) 605-610.

508

[39] M.N. Alves, P.N. Nesterenko, B. Paull, P.R. Haddad, M. Macka, Separation of superparamagnetic

509

magnetite nanoparticles by capillary zone electrophoresis using non‐complexing and complexing

510

electrolyte anions and tetramethylammonium as dispersing additive, Electrophoresis (2018).

511

[40] F. Mérida, A. Chiu-Lam, A.C. Bohórquez, L. Maldonado-Camargo, M.-E. Pérez, L. Pericchi, M. Torres-

512

Lugo, C. Rinaldi, Optimization of synthesis and peptization steps to obtain iron oxide nanoparticles with high

513

energy dissipation rates, J. Magn. Magn. Mater. 394 (2015) 361-371.

514

[41] D. Baron, P. Dolanská, Z. Medříková, R. Zbořil, J. Petr, On‐line stacking of carboxylated magnetite

515

core–shell nanoparticles in capillary electrophoresis, J. Sep. Sci. (2017).

516

[42] M. Boström, V. Deniz, G. Franks, B. Ninham, Extended DLVO theory: electrostatic and non-

517

electrostatic forces in oxide suspensions, Adv. Colloid Interface Sci. 123 (2006) 5-15.

518

[43] M. Zborowski, J.J. Chalmers, Magnetophoresis: fundamentals and applications, Wiley Encyclopedia

519

of Electrical and Electronics Engineering (1999) 1-23.

520

(25)

25

[44] G.D. Moeser, K.A. Roach, W.H. Green, T. Alan Hatton, P.E. Laibinis, High‐gradient magnetic

521

separation of coated magnetic nanoparticles, AlChE J. 50 (2004) 2835-2848.

522

[45] S. Mirshahghassemi, A.D. Ebner, B. Cai, J.R. Lead, Application of high gradient magnetic separation

523

for oil remediation using polymer-coated magnetic nanoparticles, Sep. Purif. Technol. 179 (2017) 328-334.

524

[46] P.Y. Toh, B.W. Ng, C.H. Chong, A.L. Ahmad, J.-W. Yang, C.J.C. Derek, J. Lim, Magnetophoretic

525

separation of microalgae: the role of nanoparticles and polymer binder in harvesting biofuel, RSC Advances

526

4 (2014) 4114-4121.

527

[47] P.Y. Toh, B.W. Ng, A.L. Ahmad, D.C.J. Chieh, J. Lim, Magnetophoretic separation of Chlorella sp.:

528

role of cationic polymer binder, Process Saf. Environ. 92 (2014) 515-521.

529

[48] J. Lim, S.P. Yeap, C.H. Leow, P.Y. Toh, S.C. Low, Magnetophoresis of iron oxide nanoparticles at low

530

field gradient: the role of shape anisotropy, J. Colloid Interface Sci. 421 (2014) 170-177.

531

[49] P. Zhang, S. Park, S.H. Kang, Size-dependent magnetophoresis of native single super-paramagnetic

532

nanoparticles in a microchip, Chem. Commun. 49 (2013) 7298-7300.

533

[50] R. Zhou, C. Wang, Microfluidic separation of magnetic particles with soft magnetic microstructures,

534

Microfluid. Nanofluidics 20 (2016) 48.

535

[51] V. Kumar, P. Rezai, Multiplex Inertio-Magnetic Fractionation (MIMF) of magnetic and non-

536

magnetic microparticles in a microfluidic device, Microfluid. Nanofluidics 21 (2017) 83.

537

[52] S. Dutz, M. Hayden, U. Häfeli, Fractionation of magnetic microspheres in a microfluidic spiral:

538

Interplay between magnetic and hydrodynamic forces, PLoS One 12 (2017) e0169919.

539

[53] V. Cardoso, D. Miranda, G. Botelho, G. Minas, S. Lanceros-Méndez, Highly effective clean-up of

540

magnetic nanoparticles using microfluidic technology, Sens. Actuators B Chem. 255 (2018) 2384-2391.

541

[54] R. Zhou, F. Bai, C. Wang, Magnetic separation of microparticles by shape, Lab Chip 17 (2017) 401-

542

406.

543

544

(26)

26

Table 1 – Particle composition and size, magnetic field applied, separation principle, separation time and complexity of the infrastructure used for M(N)Ps and

545

magnetically susceptible RBCs between 2013 and 2018.

546

Particle composition Particle size (nm) Magnetic field applied

Separation principle

Separation time

Complexity of the infrastructure Ref

Starch-coated magnetite (nano)particles

50–400 51 mm length

Grade N42 diametrically magnetized NdFeB cylinder

MFFF <1 hour Basic laboratory equipment [25]

Dextran-coated magnetite nano(particles)

90 and 200 Quadrupole electromagnet

MQFFF 50 min Stainless-steel cylinder within a quadrupole electromagnet implemented into a flow injection setup with downstream optical detection

[28]

RBCs 8000 Quadrupole

magnet

MQFFF 25 min Cylindrical flow channel centred inside of a quadrupole magnet with downstream light scattering detection

[29]

Carboxydextran-coated maghemite nanoparticles

6 – 60 n.a. AF4 15 min AF4 instrument connected to

MALLS detection

[32]

Hybrid polymer magnetic micelles

54 n.a. AF4 15 min AF4 instrument connected to UV

and MALLS detection

[33]

Lipid and polystyrene sulfonate-coated magnetite nanoparticles

50 and 100 n.a. CyEFFF 30 min HPLC pump connected to an EFFF

channel with downstream UV detection; ac and dc voltages induced by a signal generator and a dc power supply

[34]

Polyvinylpyrrolidone (PVP)- coated magnetic particles

127 0.56 T permanent

magnetic assembly consisting of two 2x 4 x 0.5 inch

HGMS 1 hour HGMS instrument [45]

(27)

27

NdFeB blockswith

a minimum gap of 5/8 inch

Poly (diallyl

dimethylammonium chloride) (PDDA) and chitosan (Chi)-coated magnetic nanoparticles

50 NdFeB permanent

magnet

LGMS 6 min Basic laboratory equipment [46]

Poly(diallyldimethylamonium chloride) (PDDA)-coated iron oxide nanoparticles

50 nm (spherical) and 20 x 300 nm (rod-like)

Cylindrical shaped N50-graded NdBFe (1.20 T) and Alnico permanent magnet (1.45 T) with 14 mm in diameter and 15 mm in length

LGMS 6 hours Basic laboratory equipment [48]

Bare and carboxylated iron oxide nanoparticles

7 - 13 (bare iron oxide

nanoparticles) and 10 (carboxylated iron oxide nanoparticles)

n.a. CE 12 min CE instrument [39]

Carboxylated iron oxide nanoparticles

75 n.a. CE 5 min CE instrument [41]

Polydisperse magnetic particles

150 and 500 Permanent NbFeB magnet (6 mm length, 4 mm width, and 3 mm thickness)

Microchip magnetophoresis

15 sec Microchip fabricated in PDMS with a side permanent magnet, a Gaussmeter, and DIC detection

[49]

Iron oxide magnetic beads 5000 Soft magnetic microstructures made of a mixture of iron powder and PDMS into a

Microchip magnetophoresis

<2 sec Microchip fabricated in PDMS mounted in an inverted microscope connected to a high-speed camera, placed in the centre of parallel permanent magnets

[50]

(28)

28

prefabricated

channel Polystyrene (5000 and 11000

nm) and polyethylene (35000 nm)

magnetic beads

5000, 11000, and 35000

Permanent NbFeB N42 grade cuboid magnet

Microchip magnetophoresis

15 min Microchip fabricated in PDMS with a side permanent magnet mounted in an inverted microscope

connected to a high-speed camera [51]

Microspheres composed of styrene-maleic acid copolymer matrix

encapsulating 50% by mass magnetite cores

2000, 6000 and 12000

Octupolar array of permanent magnets

Microchip magnetophoresis

10 sec Microchip fabricated in PDMS with a spiral channel centred with respect to the octupolar magnetic array

[52]

Magnetite nanoparticles 10 Permanent NbFeB

magnet

Microchip magnetophoresis

>1 hour Microchip fabricated in PDMS with permanent magnet mounted in an inverted microscope connected to a high-speed camera

[53]

Magnetite-doped and uncross-linked polystyrene particles with spherical and elliptical shapes

7000 Halbach array Microchip

magnetophoresis

<2 sec Microchip fabricated in PDMS placed in the center of the Halbach array mounted in an inverted microscope connected to a high- speed camera

[54]

AF4: Asymmetrical field flow fractionation; CE: Capillary electrophoresis; CyEFFF: Cyclical electrical field flow fractionation; HGMS: High gradient magnetic

547

separation; LGMS: Low gradient magnetic separation; MALLS: Multi-angle laser light scattering; MFFF: Magnetic field flow fractionation; MQFFF: Magnetic

548

quadrupole field flow fractionation; RBCs: Red blood cells.

549

Referanser

RELATERTE DOKUMENTER

The main goal of this study is determine the magnetic field values of chromospheric spicules from high cadence, high spatial resolution data using the weak field approximation.. A

The two sets of magnetic fluid samples, one with citric acid (MF/CA) and the other with oleic acid (MF/OA) coated magnetic nanoparticles, respectively, achieved

Figure 1.4: The main subjects of this thesis: (a) the real drilling mud, subject to the Earth’s magnetic field B; (b) the model of the drilling mud, a suspension of magnetic

(b) A paramagnetic material will orient its internal magnetic moments along the external magnetic field, whereas a diamagnetic material will create a magnetic moment in the

The LIMMCAST programme aims to model the essential features of the flow field in a continuous casting process, namely the flow field in the tundish, in the

A two-step method in- corporating the seeding of an unperturbed base flow with optimal linear perturbations in a high magnetic field strength regime shows that no increase in

Abbreviations: AF4, Asymmetric flow field flow fractionation; CLS, Centrifugal liquid sedimentation; CoV, Coefficient of variation; Dg, Geometric diameter or diameter of gyration;

This report will review a selection of the existing sensor technology in the field of magnetic, acoustic, electric, pressure, seismic and hydrodynamic (flow) sensing.. The most common